Photoferroelectric and Photopiezoelectric Properties of Organometal

Mar 22, 2016 - Close examination of the structure reveals that the Ge atom is displaced by 0.23 Å away from the center of I6 cage. The significant di...
0 downloads 7 Views 901KB Size
Subscriber access provided by BRIGHAM YOUNG UNIV

Letter

Photoferroelectric and Photopiezoelectric Properties of Organometal Halide Perovskites Shi Liu, Fan Zheng, Ilya Grinberg, and Andrew M. Rappe J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.jpclett.6b00527 • Publication Date (Web): 22 Mar 2016 Downloaded from http://pubs.acs.org on March 25, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Photoferroelectric and Photopiezoelectric Properties of Organometal Halide Perovskites Shi Liu,∗,† Fan Zheng,‡ Ilya Grinberg,‡ and Andrew M. Rappe∗,‡ †Geophysical Laboratory, Carnegie Institution for Science, Washington, D.C. 20015-1305 USA ‡The Makineni Theoretical Laboratories, Department of Chemistry, University of Pennsylvania, Philadelphia, PA 19104–6323 USA E-mail: [email protected]; [email protected]

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract Piezoelectrics play a critical role in various applications. The permanent dipole associated with the molecular cations in organometal-halide perovskites (OMHPs) may lead to spontaneous polarization and thus piezoelectricity. Here, we explore the piezoelectric properties of OMHPs with density functional theory. We find that the piezoelectric coefficient depends sensitively on the molecular ordering, and that the experimentally observed light-enhanced piezoelectricity is due to to a non-polar to polar structural transition. By comparing OMHPs with different atomic substitutions in the ABX3 architecture, we find that the displacement of the B–site cation contributes to nearly all the piezoelectric response, and that the competition between A–X hydrogen bond and B–X metal-halide bond in OMHPs controls the piezoelectric properties. These results highlight the potential of the OMHP architecture for designing new functional photoferroelectrics and photopiezoelectrics.

2

ACS Paragon Plus Environment

Page 2 of 23

Page 3 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

d33 Table of Content

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Organometal-halide perovskites (OMHPs) with their unprecedented rate of increasing power conversion efficiency (PCE) 1 have transformed photovoltaic (PV) research and development. These materials, especially methylammonium lead iodide (MAPbI3 ), are promising candidates for next-generation low cost-to-power PV technologies. 2–8 Besides the solar cell application, the OMHP architecture, consisting of ABX3 (organic monovalent cation, A; divalent metal, B; inorganic or organic anion, X), could serve as a promising platform for the design and optimization of materials with desired functionalities. By substituting ABX3 atomic constituents, it is feasible to controllably tune the structural, electronic, optical, and magnetic properties of functional materials. Piezoelectrics (materials that can convert between electrical and mechanical energy) is a class of technological important materials with a variety of applications, ranging from waveguide devices, ultrasound transducers and fuel injectors to accelerometers and gyroscopes. 9,10 The permanent dipole provided by the organic molecule in OMHPs is suggested to give rise to spontaneous polarization, which may play an important role in the photovoltaic working mechanism. 11–13 Switchable polar domains in β-MAPbI3 have been observed via piezoresponse force microscopy (PFM). 14 Though the robustness of the room-temperature ferroelectricity in MAPbI3 is still under debate, 15–18 the piezoelectric response of MAPbI3 has been demonstrated experimentally via PFM. 18 In particular, a five-fold increase of piezoelectric coefficient under illumination was observed. 18 Furthermore, the piezoelectric coefficient of MAPbI3 under light (≈25 pm/V) is comparable to inorganic ferroelectric thin films, such as lead zirconate titanate (PZT) thin films. 19 The optically-tunable high piezoelectric response in MAPbI3 suggests potential applications of OMHPs for optical piezoelectric transducers and light-driven electromechanical transistors. In this work, we investigate the piezoelectric properties of OMHPs with density functional theory (DFT). The interplay between the A-site molecular ordering and the macroscopic piezoelectric response is explored by quantifying the polarizationstrain (P –η) and polarization-stress (P –σ) relationships in two model structures with designed molecular orientations. We further study the effect of atomic substitution with

4

ACS Paragon Plus Environment

Page 4 of 23

Page 5 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

+ different organic molecules at the A site (CH3 NH+ 3 , CF3 NH3 ), metal cations at the B

site (Pb2+ , Sn2+ , Ge2+ ), and halide anions at the X site (I− , Cl− ). Our calculations show that the competition between the A–X hydrogen bond and the B–X metalhalide bond in OMHPs gives rise to a range of piezoelectric properties. The strong dependence of piezoelectric coefficients on the molecular ordering suggests that the light-enhanced piezoelectricity is likely due to the light-induced molecular reorientation under external voltage. This work demonstrates that the OMHP architecture could serve as a promising platform for designing functional piezoelectrics with robust optical tunability. All calculations are performed with a plane-wave implementation of the PerdewBurke-Ernzerhof (PBE) density functional 20 in the QUANTUM-ESPRESSO code. 21 The optimized norm-conserving pseudopotentials 22,23 are generated using the OPIUM 22,23 code with the following electronic states included as valence states: Pb: 5d, 6s, 6p; I: 4d, 5s, 5p; C: 2s, 2p; N: 2s, 2p; H: 1s. A 4×4×4 Monkhorst-Pack grid is used, consistent with our previous work. 12,24 The effect of molecular orientation on piezoelectricity is explored by calculating the piezoelectric coefficients of two model structures, namely M1 with non-zero total polarization and M2 with net-zero polarization, shown in Figure 1. The M1 structure has an isotropic orientation of the MA cations in the ab-plane, and an anisotropic orientation with respect to the c-axis, giving rise to a net polarization along the c-axis. The molecular dipoles in the M2 structure adopt an antiferroelectric ordering and the structure is globally (nearly) paraelectric. These two models are representatives of possible molecular orientations in OMHPs and have been used previously to reveal the effect of polar order on bulk photovoltaic effects 24 and the dynamics of molecular cations. 25 First-principles techniques to determine piezoelectric coefficients have been widely applied to inorganic materials. 26–29 As piezoelectricity is a fundamental process incorporating different electromechanical effects, multiple piezoelectric equations exist. 10,29,30 The direct piezoelectric effect that links the induced polarization that develops in direction i (∆Pi ) with an applied stress with component j (σj in Voigt notation)

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 23

can be described by

∆Pi = dij σj

(1)

where {dij } is a third-rank tensor and each dij parameter is usually referred to as a direct piezoelectric coefficient or piezoelectric strain coefficient, in units of pC/N. Another piezoelectric equation connecting the strain η and the polarization is given by: ∆Pi = eij ηj

(2)

where the eij parameter is called the piezoelectric stress coefficient in the unit of C/m2 . It is noted that the eij and dij parameters are related to each other via elastic compliances and/or stiffness, though dij is easier to measure experimentally. In this study, we focus on the evaluation of d33 and e33 (piezoelectric response along the c axis). The total induced polarization along the c axis can be expressed as ∆P3 = e33 η3 +e31 (η1 +η2 ), where η1 = (a − a0 )/a0 , η2 = (b − b0 )/b0 and η3 = (c − c0 )/c0 are strains along the a, b, and c axis, and a0 , b0 , and c0 are lattice constants of an unstrained structure. When only applying strain along the c axis, e33 can be evaluated from the slope of polarization vs. strain (η3 ) curve. For a given strain η3 , the internal atomic coordinates are fully relaxed until the residual forces on atoms are less than 1.0×10−5 Ryd/Bohr, and the polarization of the optimized structure is determined with the Berry’s phase approach. The estimation of d33 with DFT is less straightforward, due to the difficulty of defining the stress tensor. We adopt the finite difference method developed in Ref. 31. After applying a strain along the [001] direction, the other lattice parameters (lattice constants a and b, and lattice angles) as well as internal atomic coordinates are fully optimized until all stress tensor elements except σ3 are smaller than 1.0×10−3 GPa. The polarization values of various strained structures are then calculated, and the slope of P vs. σ3 determines the value of d33 . For each material investigated, the structure is first fully optimized to obtain the equilibrium lattice constants, and ±0.5% strains are applied to the ground-state structures to obtain linear response.

6

ACS Paragon Plus Environment

Page 7 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

A previous first-principles benchmark study highlighted the importance of dispersive interactions and the plane-wave cutoff energy (Ec ) in accurately determining structural properties of OMHPs. 32 Because the value of piezoelectric coefficient depends sensitively on the lattice constants, we first compare the optimized structural parameters of MAPbI3 obtained with four different methods: pure PBE density functional with Ec = 50 Ry, PBE with Ec = 80 Ry, PBE plus Grimme dispersion correction 33 (PBED2) with Ec = 50 Ry, and PBE-D2 with Ec = 80 Ry. As shown in Table 1, the value of unit cell volume predicted by PBE-D2 with Ec = 80 Ry best agrees with the experimental results. Therefore, we choose this method for the following piezoelectric coefficient calculations. Figure 2 shows the P -σ3 and P -η3 curves for MAPbI3 in polar M1 and non-polar M2 configurations. The computed piezoelectric coefficients of M1 are d33 = 31.4 pm/V and e33 = 0.83 C/m2 , dramatically higher than those of M2, d33 = 3.1 pm/V and e33 = 0.07 C/m2 . Note that the values of d33 of M1 and M2 agree reasonably well with experimental values 18 measured under white light (25 pm/V) and dark (6 pm/V), respectively. It is suggested that illumination could weaken the hydrogen bond between the MA cation and the inorganic Pb-I scaffold, thus effectively reducing the rotational barrier of the MA cation. 34 This allows the formation of a more polar structure with aligned MA cations in the presence of an electric field, which could arise from the photovoltage across the thin film. Therefore, the light-enhanced piezoelectric response is likely due to the transformation of a configuration with less-ordered molecular orientation (e.g., M2–like) to a configuration with more-ordered molecular orientation (e.g., M1–like). To reveal the origin of piezoelectricity, we decompose the total polarization into the contributions from the A-site MA cations and B-site Pb atoms in M1. The polarization ∗ (Pb) × D(Pb)/V , where V is the volume of the from Pb is estimated with PPb = Z33 u u ∗ (Pb) is the Born primitive unit cell (1/4 of the volume of the tetragonal supercell), Z33

effective charge of Pb and D(Pb) is the average displacement of Pb along the c axis ∗ (Pb) is +4.24, with respect to the center of its I6 cage. The calculated value of Z33

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

significantly larger than the nominal charge of Pb (+2.0) in a pure ionic picture. This indicates the strong covalency of the Pb–I bonds and the presence of dynamic charge transfer coupled with the change of Pb-I bond length. The polarization from MA+ molecules is approximated with PMA = µ cos(θ)/Vu , where µ is the dipole of the MA+ molecule and θ is the average angle between the molecular dipole and the c axis. The site-resolved polarization and d33 are presented in Figure 3a. It is clear that though both Pb displacements and molecular dipoles are responsible for the total polarization, Pb atoms contribute nearly all the piezoelectric response, with the contribution from the MA+ molecules negligible. As shown in Figure 3b, the angle between the MA+ and the c axis remains nearly constant, while the atomic displacement of Pb changes linearly with the stress. We further explore the effect of chemical substitution on the piezoelectric properties of OMHPs. Table 2 presents the values of d33 and e33 for MAPbI3 , MASnI3 , MAGeI3 , CF3 NH3 PbI3 , MAPbCl3 and MASnCl3 . For all the materials investigated, the M2 configuration has lower polarization and smaller d33 and e33 than the M1 configuration. It is noted, however, that MAGeI3 of M2 configuration still has a moderate polarization of 8.92µC/cm2 despite the molecular dipoles nearly canceling each other out. Close examination of the structure reveals that the Ge atom is displaced by 0.23 ˚ A away from the center of GeI6 cage. The significant displacement is due to the small size of Ge relative to the large A cations and I6 octahedron. As shown in Figure 4, for MAPbI3 , MASnI3 and MAGeI3 , three OMHPs differing only by the B-site element, MAGeI3 M1 has the highest polarization, while MASnI3 M1 has the highest piezoelectric coefficient (d33 = 100.9 pC/N), comparable to classic inorganic ferroelectric PbTiO3 (d33 = 80 pC/N). 31 The higher d33 in MASnI3 is attributed to the larger Born effective charge ∗ (Sn) = +4.64, compared to Z ∗ (Ge) = +3.60 and Z ∗ (Pb) = +4.24. Comparing Z33 33 33

MAPbCl3 and MASnCl3 to MAPbI3 and MASnI3 , we find that substituting I with Cl dramatically reduces d33 and e33 , which can be explained by the reduced covalency (more ionic character) of Pb(Sn)-Cl bonds. Born effective charge calculations show ∗ of Sn is reduced to +3.39 in MASnCl . This means a smaller charge transfer, that Z33 3

8

ACS Paragon Plus Environment

Page 8 of 23

Page 9 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

thus less polarization change, will occur upon change of the Sn-Cl bond length, causing smaller piezoelectric response.

As demonstrated with CF3 NH3 PbI3 , replacing the MA+ with a more polar molecular cation greatly increases both the polarization and the piezoelectric response. The larger molecular dipoles not only contribute directly to the polarization, but also increase the hydrogen bonding strength between NH3 and I, which indirectly weakens the bonding between I and Pb. This is supported by the larger Pb displacement (≈ 0.30 ˚ A) at zero stress condition compared to that of ≈ 0.09 ˚ A in MAPbI3 . In other words, the Pb atoms in CF3 NH3 PbI3 have a softer potential energy surface and therefore a stronger response to external perturbation, resulting in higher piezoelectricity. Many processes may contribute to the light-induced molecular dipole reorientation. Here, we discuss possible mechanisms for the light-enhanced piezoelectricity in more detail. It is suggested that the light illumination, by weakening the hydrogen bond between the MA cation and the inorganic PbI3 sublattice, could make the molecular cations easier to rotate and to align with external field. This is supported by DFT calculations which reveal a reduced binding energy of the MA cation to the inorganic cage at the triplet state. 34 Additionally, as the band gap of MAPbI3 is in the visible light region, a significant amount of above band gap excitation occurs, with the excess energy dissipating as heat (kinetic energy of molecules) when the hot carriers relax to the band edges. We have recently shown that the molecules play an important role in carrier relaxation, with MA translation, CH/NH twisting, and CH/NH stretching modes assisting the process. 35 We thus develop a simple analytical model here to examine the magnitude of light-induced thermalization in a MAPbI3 thin film. 36 Shown in Figure 5a is the absorption spectrum α(E) computed from the DFT optical dielectric function. The dielectric function is summed over the Brillouin zone and enough empty bands to converge. Assuming that all the energy released from hot carrier relaxation is transformed into kinetic energy (κ) of molecules, the average thermal energy received by each molecule at depth D per second can be estimated as:

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

κ (D) = A

Z ∞ Eg

dE

Page 10 of 23

 I0 (E) e−α(E)D (E − Eg )  1 − e−α(E)∆d E

(3)

where I0 (E) is the sunlight spectrum, ∆d is the thickness of a single layer of molecules (equal to the lattice constant of MAPbI3 , 0.63 nm), A is the area per molecule, and Eg is the band gap of 1.65 eV calculated from DFT. As plotted in Figure 5b, the energy obtained by a single molecule per second has a strong dependence on the depth due to the exponential decay of the light intensity. In particular, the molecules near the top surface of the sample can obtain significant energy (much larger than room temperature energy scale) to overcome the barrier and re-orient their dipole directions under the applied voltage, which is likely to result in a more polar structure with large piezoelectricity, as measured in the experiments. The reorientation of the molecular cations should occur rapidly (∼ps timescale), 25,37 however, the adjustment of the inorganic Pb–I scaffold could be much slower. 38 Juarez-Perez et al.

found that the

low-frequency (< 104 Hz) dielectric constant of MAPbI3 is enhanced significantly upon light illumination, 39 which suggests a structural variation occurring at a timescale of 10−4 s. Recent experiments also find that the structural transformation in MAPbI3 between the dark and the light is reversible. 38 It is also noted that as the lattice constants of MAPbI3 change under temperature, the piezoelectric constants are expected to be strongly temperature dependent. We also used the experimental absorption from Ref. 40 to compute the heat transfer, which shows very close value (11.2 meV) compared to DFT result (11.8 meV) for D = 50 nm. The DFT absorption spectrum overestimates the absorption in the high-energy range but slightly underestimates the low energy range. The overall heat transfer to the molecule will not be affected significantly. In summary, the piezoelectric properties of several OMHPs are studied with ab initio density functional theory. For MAPbI3 , the calculated piezoelectric coefficients of the polar and non-polar configurations agree reasonably well with experimental values measured under white light and dark, respectively. This suggests a light-driven molecular reordering, which could be attributed to the weakened hydrogen bonds between

10

ACS Paragon Plus Environment

Page 11 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

MA cations and inorganic cages in the excited state and the thermalization arising from hot carrier relaxation. The photopiezoelectricity of OMHPs offers a potential avenue to optical transducers. Comparing the piezoelectric properties of OMHPs with different organic molecules, metal cations and halide anions helps to identify several design principles for tuning the piezoelectric coefficients. We find that the atomic displacement of B–site metal cations is responsible for the piezoelectric response, and therefore creating a softer energy profile vs. cation displacement is critical for enhancing d33 and e33 . Our finding highlights the potential of OMHPs for the design and optimization of functional photopiezoelectrics and photoferroelectrics.

Acknowledgements S.L. is supported by the Carnegie Institution for Science and the Office of Naval Research. F.Z. acknowledges the support of the Office of Naval Research, under grant N00014-14-1-0761. I.G. acknowledges the support of the Office of Naval Research, under grant N00014-12-1-1033. A.M.R. acknowledges the support of the Department of Energy Office of Basic Energy Sciences, under grant DE-FG02-07ER46431. All the authors thank the National Energy Research Scientific Computing Center of the Department of Energy and the High-Performance Computing Modernization Office of the Department of Defense.

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 23

Table 1: Structural parameters of tetragonal MAPbI3 in the M1 configuration, compared to experimental data.

M1

M2

Method PBE PBE PBE-D2 PBE-D2 PBE PBE PBE-D2 PBE-D2 experiment

Ec (Ry) 50 80 50 80 50 80 50 80 Ref 8 Ref 41 Ref 3

a (˚ A) 8.91 8.97 8.64 8.67 8.88 8.99 8.63 8.69 8.85 8.86 8.85

b (˚ A) 8.81 8.92 8.68 8.74 8.89 8.96 8.64 8.69 8.85 8.86 8.85

12

c (˚ A) 13.09 13.29 12.77 12.97 13.06 13.22 12.82 12.97 12.64 12.66 12.44

α 90.00 90.01 90.05 90.07 91.18 91.27 90.41 90.09 90.0 90.0 90.0

ACS Paragon Plus Environment

β 90.00 90.03 89.86 89.87 91.39 90.04 90.45 90.10 90.0 90.0 90.0

γ 90.03 90.00 89.92 90.00 89.75 89.71 90.17 90.06 90.0 90.0 90.0

V (˚ A3 ) 1027.10 1062.87 957.93 981.90 1030.33 1063.54 955.71 979.83 990.00 993.80 974.33

Page 13 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Table 2: Ground state polarization along the c axis (Pc ) in µC/cm2 , piezoelectric coefficients d33 in pC/N and e33 in C/m2 for various OMHPs of M1 and M2 configurations

OMHP MAPbI3 MASnI3 MAGeI3 MAPbCl3 MASnCl3 CF3 NH3 PbI3

Pc 5.59 7.75 13.68 5.96 5.46 15.8

M1 d33 e33 31.4 0.83 100.9 1.85 27.4 0.49 6.7 0.16 4.1 0.05 248 1.59

13

Pc 0.07 -0.22 8.92 -0.01 0.01 -0.09

ACS Paragon Plus Environment

M2 d33 3.1 23.7 32.4 – – 4.23

e33 0.07 0.51 0.52 – – 0.16

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References (1) NREL. Research Cell Efficiency Records. http://www.nrel.gov/ncpv/images/ efficiency_chart.jpg (accessed Mar. 9, 2016). (2) Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites As Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. (3) Baikie, T.; Fang, Y.; Kadro, J. M.; Schreyer, M.; Wei, F.; Mhaisalkar, S. G.; Graetzel, M.; White, T. J. Synthesis and Crystal Chemistry of the Hybrid Perovskite (CH3 NH3 )PbI3 for Solid-State Sensitised Solar Cell Applications. J. of Mater. Chem. A 2013, 1, 5628–5641. (4) Im, J.-H.; Lee, C.-R.; Lee, J.-W.; Park, S.-W.; Park, N.-G. 6.5% Efficient Perovskite Quantum-Dot-Sensitized Solar Cell. Nanoscale 2011, 3, 4088–4093. (5) Filip, M. R.; Eperon, G. E.; Snaith, H. J.; Giustino, F. Steric Engineering of Metal-Halide Perovskites with Tunable Optical Band Gaps. Nat. Commun. 2014, 5, 5757–1–9. (6) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J. Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, 643–647. (7) Etgar, L.; Gao, P.; Xue, Z.; Peng, Q.; Chandiran, A. K.; Liu, B.; Nazeeruddin, M. K.; Graetzel, M. Mesoscopic CH3 NH3 PbI3 /TiO2 Heterojunction Solar Cells. J. Am. Chem. Soc. 2012, 134, 17396–17399. (8) Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G. Semiconducting Tin and Lead Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties. Inorg. Chem. 2013, 52, 9019–9038.

14

ACS Paragon Plus Environment

Page 14 of 23

Page 15 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(9) Jaffe, B.; Cook, W.; Jaffe, H. Piezoelectric ceramics; Academic Press, New York, 1971. (10) Ballato, A. Piezoelectricity: Old effect, New thrusts. IEEE Transactions on Ultrasonics, Ferroelectrics and Frequency Control 1995, 42, 916–926. (11) Frost, J. M.; Butler, K. T.; Brivio, F.; Hendon, C. H.; van Schilfgaarde, M.; Walsh, A. Atomistic Origins of High-Performance in Hybrid Halide Perovskite Solar Cells. Nano Lett. 2014, 14, 2584–2590. (12) Liu, S.; Zheng, F.; Koocher, N. Z.; Takenaka, H.; Wang, F.; Rappe, A. M. Ferroelectric Domain Wall Induced Band Gap Reduction and Charge Separation in Organometal Halide Perovskites. J. Phys. Chem. Lett. 2015, 6, 693–699. (13) Ma, J.; Wang, L.-W. Nanoscale Charge Localization Induced by Random Orientations of Organic Molecules in Hybrid Perovskite CH3 NH3 PbI3 . Nano Lett. 2015, 15, 248–253. (14) Kutes, Y.; Ye, L.; Zhou, Y.; Pang, S.; Huey, B. D.; Padture, N. P. Direct Observation of Ferroelectric Domains in Solution-Processed CH3 NH3 PbI3 Perovskite Thin Films. J. Phys. Chem. Lett. 2014, 5, 3335–3339. (15) Xiao, Z.; Yuan, Y.; Shao, Y.; Wang, Q.; Dong, Q.; Bi, C.; Sharma, P.; Gruverman, A.; Huang, J. Giant Switchable Photovoltaic Effect in Organometal Trihalide Perovskite Devices. Nat. Mater. 2015, 14, 193–198. (16) Fan, Z.; Xiao, J.; Sun, K.; Chen, L.; Hu, Y.; Ouyang, J.; Ong, K. P.; Zeng, K.; Wang, J. Ferroelectricity of CH3 NH3 PbI3 Perovskite. J. Phys. Chem. Lett. 2015, 6, 1155–1161. (17) Kim, H.-S.; Kim, S. K.; Kim, B. J.; Shin, K.-S.; Gupta, M. K.; Jung, H. S.; Kim, S.-W.; Park, N.-G. Ferroelectric Polarization in CH3 NH3 PbI3 Perovskite. J. Phys. Chem. Lett. 2015, 6, 1729–1735.

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(18) Coll, M.; Gomez, A.; Mas-Marza, E.; Almora, O.; Garcia-Belmonte, G.; CampoyQuiles, M.; Bisquert, J. Polarization Switching and Light-Enhanced Piezoelectricity in Lead Halide Perovskites. J. Phys. Chem. Lett. 2015, 6, 1408–1413. (19) Lefki, K.; Dormans, G. J. M. Measurement of Piezoelectric Coefficients of Ferroelectric Thin Films. J. Appl. Phys. 1994, 76, 1764. (20) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. (21) Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G. L.; Cococcioni, M.; Dabo, I. et al. Quantum ESPRESSO: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys.: Condens. Matter 2009, 21, 395502–20. (22) Rappe, A. M.; Rabe, K. M.; Kaxiras, E.; Joannopoulos, J. D. Optimized Pseudopotentials. Phys. Rev. B Rapid Comm. 1990, 41, 1227–1230. (23) Ramer, N. J.; Rappe, A. M. Designed Nonlocal Pseudopotentials for Enhanced Transferability. Phys. Rev. B 1999, 59, 12471–12478. (24) Zheng, F.; Takenaka, H.; Wang, F.; Koocher, N. Z.; Rappe, A. M. First-Principles Calculation of Bulk Photovoltaic Effect in CH3 NH3 PbI3 and CH3 NH3 PbI3−x Clx . J. Phys. Chem. Lett. 2015, 6, 31–37. (25) Quarti, C.; Mosconi, E.; Angelis, F. D. Structural and Electronic Properties of Organo-Halide Hybrid Perovskites from Ab Initio Molecular Dynamics. Phys. Chem. Chem. Phys. 2015, 17, 9394–9409. (26) S´ aghi-Szab´ o, G.; Cohen, R. E.; Krakauer, H. First-Principles Study of Piezoelectricity in Tetragonal PbTiO3 and PbZr1/2 Ti1/2 O3 . Phys. Rev. B 1999, 59, 12771–12776. (27) S´ aghi-Szab´ o, G.; Cohen, R. E.; Krakauer, H. First-Principles Study of Piezoelectricity in PbTiO3 . Phys. Rev. Lett. 1998, 80, 4321.

16

ACS Paragon Plus Environment

Page 16 of 23

Page 17 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(28) Bernardini, F.; Fiorentini, V. First-Principles Calculation of the Piezoelectric Tensor d of III–V Nitrides. Appl. Phys. Lett. 2002, 80, 4145. (29) Bellaiche, L. Piezoelectricity of Ferroelectric Perovskites from First Principles. Current Opinion in Solid State and Materials Science 2002, 6, 19–25. (30) Nye, J. Physical Properties of Crystals: Their Representation by Tensors and Matrices; Oxford science publications; Clarendon Press, 1985. (31) Shi, J.; Grinberg, I.; Wang, X.; Rappe, A. M. Atomic Sublattice Decomposition of Piezoelectric Response in Tetragonal PbTiO3 , BaTiO3 , and KNbO3 . Phys. Rev. B 2014, 89, 094105. (32) Egger, D. A.; Kronik, L. Role of Dispersive Interactions in Determining Structural Properties of Organic–Inorganic Halide Perovskites: Insights from FirstPrinciples Calculations. J. Phys. Chem. Lett. 2014, 5, 2728–2733. (33) Grimme, S. Semiempirical GGA-type Density Functional Constructed With A Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. (34) Gottesman, R.; Haltzi, E.; Gouda, L.; Tirosh, S.; Bouhadana, Y.; Zaban, A.; Mosconi, E.; De Angelis, F. Extremely Slow Photoconductivity Response of CH3 NH3 PbI3 Perovskites Suggesting Structural Changes Under Working Conditions. The Journal of Physical Chemistry Letters 2014, 5, 2662–2669. (35) Zheng, F.; Tan, L. Z.; Liu, S.; Rappe, A. M. Rashba Spin–Orbit Coupling Enhanced Carrier Lifetime in CH 3 NH 3 PbI 3. Nano Letters 2015, 15, 7794–7800. (36) Li, L.; Wang, F.; Wu, X.; Yu, H.; Zhou, S.; Zhao, N. Carrier-Activated Polarization in Organometal Halide Perovskites. J. Phys. Chem. C 2016, 120, 2536–2541. (37) Knop, O.; Wasylishen, R. E.; White, M. A.; Cameron, T. S.; Oort, M. J. M. V. Alkylammonium Lead Halides. Part 2. CH3 NH3 PbX3 (X = Cl, Br, I) Perovskites: Cuboctahedral Halide Cages With Isotropic Cation Reorientation. Canadian Journal of Chemistry 1990, 68, 412–422.

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(38) Gottesman, R.; Gouda, L.; Kalanoor, B. S.; Haltzi, E.; Tirosh, S.; RoshHodesh, E.; Tischler, Y.; Zaban, A.; Quarti, C.; Mosconi, E. et al. Photoinduced Reversible Structural Transformations in Free-Standing CH3 NH3 PbI3 Perovskite Films. J. Phys. Chem. Lett. 2015, 6, 2332–2338. (39) Juarez-Perez, E. J.; Sanchez, R. S.; Badia, L.; Garcia-Belmonte, G.; Kang, Y. S.; Mora-Sero, I.; Bisquert, J. Photoinduced Giant Dielectric Constant in Lead Halide Perovskite Solar Cells. J. Phys. Chem. Lett. 2014, 5, 2390–2394. (40) Park, B.-W.; Jain, S. M.; Zhang, X.; Hagfeldt, A.; Boschloo, G.; Edvinsson, T. Resonance Raman and Excitation Energy Dependent Charge Transfer Mechanism in Halide-Substituted Hybrid Perovskite Solar Cells. ACS Nano 2015, 9, 2088– 2101. (41) Poglitsch, A.; Weber, D. Dynamic Disorder in Methylammoniumtrihalogenoplumbates (II) Observed by Millimeter-Wave Spectroscopy. J. Chem. Phys. 1987, 87, 6373–6378.

18

ACS Paragon Plus Environment

Page 18 of 23

Page 19 of 23

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

(a) 6.2

(b)6.2 M1: e33 = 0.83 C/m2

Polarization (µC/cm2 )

Polarization (µC/cm2 )

M1: d33 = 31.4 pm/V

6.0 5.8 5.6 5.4 5.2 5.0 1.5

(c)

1.0

0.5 0.0 0.5 Stress (kbar)

1.0

5.8 5.6 5.4 5.2

(d) 0.14

M2: d33 = 3.0 pm/V

Polarization (µC/cm2 )

0.12

6.0

5.0 0.994 0.996 0.998 1.000 1.002 1.004 1.006

1.5

0.14 Polarization (µC/cm2 )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

0.10 0.08 0.06 0.04 0.02 0.00 2.0 1.5 1.0 0.5 0.0 0.5 1.0 1.5 2.0 Stress (kbar)

0.12

M2: e33 = 0.07 C/m2

0.10 0.08 0.06 0.04 0.02 0.00 0.994 0.996 0.998 1.000 1.002 1.004 1.006

Figure 2: Piezoelectric properties of MAPbI3 in M1 (a-b) and M2 (c-d) configurations. (a,c) Polarization as a function of stress. (b,d) Polarization as a function of strain.

20

ACS Paragon Plus Environment

Page 20 of 23

(a)

3.0

0.12

(b)

80

Pb: d33 = 35.7 pm/V Atomic Displacement (Angstrom)

MA: d33 = 0.2 pm/V

2.8 Polarization (µC/cm2 )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

2.6

2.4

2.2

2.0

1.8 1.5

1.0

0.5

0.0 Stress (kbar)

0.5

1.0

0.11

78

0.10 76 0.09 74 0.08 72

0.07

0.06 1.5

1.5

D (Pb) (MA)

Angle (Degree)

Page 21 of 23

1.0

0.5

0.0 Stress (kbar)

0.5

1.0

70 1.5

Figure 3: (a) Site-resolved polarization as a function of stress. (b) Atomic displacement and molecular orientational angle as a function of stress.

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

25 MAPbI3 : d33 = 31.4 pm/V MASnI3 : d33 = 86.2 pm/V MAGeI3 : d33 = 27.4 pm/V MAPbCl3 : d33 = 6.7 pm/V

20 Polarization (µC/cm2 )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 23

MASnCl3 : d33 = 4.1 pm/V CF3 NH3 PbI3 : d33 = 247.6 pm/V

15

10

5

0 2.0

1.5

1.0

0.5 0.0 0.5 Stress (kbar)

1.0

1.5

2.0

Figure 4: Stress-dependence of polarization for various OMHPs, all with M1 molecular orientation.

22

ACS Paragon Plus Environment

Page 23 of 23

b)

4

2

0

1

2

3

4

Heat transformed (meV)

a) 6 (1/m)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

50 40 30 20 10

0

100

200

300

400

Depth D (nm)

Photon Energy (eV)

Figure 5: (a) The DFT calculated absorption spectrum. (b) Estimation of the averaged energy transferred from the photo-excited carriers to each molecule at different depths (D) per second. The insert illustrates the model described by Equation 3.

23

ACS Paragon Plus Environment